• No results found

A robust self learning method for fully unsupervised cross lingual mappings of word embeddings

N/A
N/A
Protected

Academic year: 2020

Share "A robust self learning method for fully unsupervised cross lingual mappings of word embeddings"

Copied!
10
0
0

Loading.... (view fulltext now)

Full text

(1)

789

A robust self-learning method for

fully unsupervised cross-lingual mappings of word embeddings

Mikel Artetxe and Gorka Labaka and Eneko Agirre

IXA NLP Group

University of the Basque Country (UPV/EHU)

{mikel.artetxe,gorka.labaka,e.agirre}@ehu.eus

Abstract

Recent work has managed to learn cross-lingual word embeddings without parallel data by mapping monolingual embeddings to a shared space through adversarial train-ing. However, their evaluation has focused on favorable conditions, using comparable corpora or closely-related languages, and we show that they often fail in more re-alistic scenarios. This work proposes an alternative approach based on a fully un-supervised initialization that explicitly ex-ploits the structural similarity of the em-beddings, and a robust self-learning algo-rithm that iteratively improves this solu-tion. Our method succeeds in all tested scenarios and obtains the best published results in standard datasets, even surpass-ing previous supervised systems. Our implementation is released as an open source project at https://github. com/artetxem/vecmap.

1 Introduction

Cross-lingual embedding mappings have shown to be an effective way to learn bilingual word em-beddings (Mikolov et al., 2013;Lazaridou et al., 2015). The underlying idea is to independently train the embeddings in different languages us-ing monolus-ingual corpora, and then map them to a shared space through a linear transformation. This allows to learn high-quality cross-lingual rep-resentations without expensive supervision, open-ing new research avenues like unsupervised neural machine translation (Artetxe et al.,2018b;Lample et al.,2018).

While most embedding mapping methods rely on a small seed dictionary, adversarial training has recently produced exciting results in fully

unsu-pervised settings (Zhang et al.,2017a,b;Conneau et al., 2018). However, their evaluation has fo-cused on particularly favorable conditions, lim-ited to closely-related languages or comparable Wikipedia corpora. When tested on more realis-tic scenarios, we find that they often fail to pro-duce meaningful results. For instance, none of the existing methods works in the standard English-Finnish dataset fromArtetxe et al.(2017), obtain-ing translation accuracies below 2% in all cases (see Section5).

On another strand of work,Artetxe et al.(2017) showed that an iterative self-learning method is able to bootstrap a high quality mapping from very small seed dictionaries (as little as 25 pairs of words). However, their analysis reveals that the self-learning method gets stuck in poor local op-tima when the initial solution is not good enough, thus failing for smaller training dictionaries.

(2)

EN − two IT − due (two) IT − cane (dog)

0 100 200 300

[image:2.595.98.502.66.196.2]

−0.02 −0.01 0.00 0.01 0.02 −0.02 −0.01 0.00 0.01 0.02 −0.02 −0.01 0.00 0.01 0.02

Figure 1: Motivating example for our unsupervised initialization method, showing the similarity distri-butions of three words (corresponding to the smoothed density estimates from the normalized square root of the similarity matrices as defined in Section3.2). Equivalent translations (two and due) have more similar distributions than non-related words (two andcane- meaning dog). This observation is used to build an initial solution that is later improved through self-learning.

2 Related work

Cross-lingual embedding mapping methods work by independently training word embeddings in two languages, and then mapping them to a shared space using a linear transformation.

Most of these methods aresupervised, and use a bilingual dictionary of a few thousand entries to learn the mapping. Existing approaches can be classified into regression methods, which map the embeddings in one language using a least-squares objective (Mikolov et al., 2013; Shigeto et al.,2015;Dinu et al.,2015), canonical methods, which map the embeddings in both languages to a shared space using canonical correlation analy-sis and extensions of it (Faruqui and Dyer,2014; Lu et al.,2015), orthogonal methods, which map the embeddings in one or both languages under the constraint of the transformation being orthog-onal (Xing et al.,2015;Artetxe et al.,2016;Zhang et al.,2016;Smith et al.,2017), and margin meth-ods, which map the embeddings in one language to maximize the margin between the correct trans-lations and the rest of the candidates (Lazaridou et al., 2015). Artetxe et al. (2018a) showed that many of them could be generalized as part of a multi-step framework of linear transformations.

A related research line is to adapt these methods to thesemi-supervisedscenario, where the train-ing dictionary is much smaller and used as part of a bootstrapping process. While similar ideas where already explored for traditional count-based vec-tor space models (Peirsman and Pad´o,2010;Vuli´c and Moens, 2013), Artetxe et al. (2017) brought this approach to pre-trained low-dimensional word

embeddings, which are more widely used nowa-days. More concretely, they proposed a self-learning approach that alternates the mapping and dictionary induction steps iteratively, obtaining re-sults that are comparable to those of supervised methods when starting with only 25 word pairs.

(3)

embed-dings. Despite promising, they conclude that their model “is not competitive with other cross-lingual representation approaches”. Zhang et al. (2017a) use a very similar architecture, but incorporate ad-ditional techniques like noise injection to aid train-ing and report competitive results on biltrain-ingual lex-icon extraction. Conneau et al. (2018) drop the reconstruction component, regularize the mapping to be orthogonal, and incorporate an iterative re-finement process akin to self-learning, reporting very strong results on a large bilingual lexicon extraction dataset. Finally, Zhang et al. (2017b) adopt the earth mover’s distance for training, opti-mized through a Wasserstein generative adversar-ial network followed by an alternating optimiza-tion procedure. However, all this previous work used comparable Wikipedia corpora in most ex-periments and, as shown in Section 5, face diffi-culties in more challenging settings.

3 Proposed method

LetX andZ be the word embedding matrices in two languages, so that their ith row Xi∗ andZi∗ denote the embeddings of theith word in their re-spective vocabularies. Our goal is to learn the lin-ear transformation matrices WX and WZ so the mapped embeddingsXWX and ZWZ are in the same cross-lingual space. At the same time, we aim to build a dictionary between both languages, encoded as a sparse matrix Dwhere Dij = 1if thejth word in the target language is a translation of theith word in the source language.

Our proposed method consists of four sequen-tial steps: a pre-processing that normalizes the embeddings (§3.1), a fully unsupervised initializa-tion scheme that creates an initial soluinitializa-tion (§3.2), a robust self-learning procedure that iteratively im-proves this solution (§3.3), and a final refinement step that further improves the resulting mapping through symmetric re-weighting (§3.4).

3.1 Embedding normalization

Our method starts with a pre-processing that length normalizes the embeddings, then mean centers each dimension, and then length normal-izes them again. The first two steps have been shown to be beneficial in previous work (Artetxe et al., 2016), while the second length normaliza-tion guarantees the final embeddings to have a unit length. As a result, the dot product of any two embeddings is equivalent to their cosine similarity

and directly related to their Euclidean distance1, and can be taken as a measure of their similarity.

3.2 Fully unsupervised initialization

The underlying difficulty of the mapping problem in its unsupervised variant is that the word embed-ding matricesXandZ are unaligned across both axes: neither theith vocabulary itemXi∗ andZi∗ nor thejth dimension of the embeddingsX∗j and

Z∗j are aligned, so there is no direct correspon-dence between both languages. In order to over-come this challenge and build an initial solution, we propose to first construct two alternative repre-sentationsX′ andZ′ that are aligned across their

jth dimension Xj and Zj, which can later be used to build an initial dictionary that aligns their respective vocabularies.

Our approach is based on a simple idea: while the axes of the original embeddingsX andZ are different in nature, both axes of their correspond-ing similarity matrices MX = XXT andMZ =

ZZT correspond to words, which can be exploited to reduce the mismatch to a single axis. More con-cretely, assuming that the embedding spaces are perfectly isometric, the similarity matrices MX andMZ would be equivalent up to a permutation of their rows and columns, where the permutation in question defines the dictionary across both lan-guages. In practice, the isometry requirement will not hold exactly, but it can be assumed to hold ap-proximately, as the very same problem of map-ping two embedding spaces without supervision would otherwise be hopeless. Based on that, one could try every possible permutation of row and column indices to find the best match betweenMX andMZ, but the resulting combinatorial explosion makes this approach intractable.

In order to overcome this problem, we pro-pose to first sort the values in each row of MX and MZ, resulting in matrices sorted(MX) and

sorted(MZ)2. Under the strict isometry condition, equivalent words would get the exact same vec-tor across languages, and thus, given a word and its row in sorted(MX), one could apply nearest neighbor retrieval over the rows ofsorted(MZ)to find its corresponding translation.

On a final note, given the singular value de-composition X = U SVT, the similarity matrix

1

Given two length normalized vectorsuandv, u·v = cos(u, v) = 1− ||u−v||2/2.

2Note that the values in each row are sorted independently

(4)

is MX = U S2UT. As such, its square root √

MX = U SUT is closer in nature to the origi-nal embeddings, and we also find it to work better in practice. We thus computesorted(√MX)and

sorted(√MZ)and normalize them as described in Section3.1, yielding the two matricesX′ andZ′

that are later used to build the initial solution for self-learning (see Section3.3).

In practice, the isometry assumption is strong enough so the above procedure captures some cross-lingual signal. In our English-Italian exper-iments, the average cosine similarity across the gold standard translation pairs is 0.009 for a ran-dom solution, 0.582 for the optimal supervised so-lution, and 0.112 for the mapping resulting from this initialization. While the latter is far from be-ing useful on its own (the accuracy of the resultbe-ing dictionary is only 0.52%), it is substantially better than chance, and it works well as an initial solution for the self-learning method described next.

3.3 Robust self-learning

Previous work has shown that self-learning can learn high-quality bilingual embedding mappings starting with as little as 25 word pairs (Artetxe et al., 2017). In this method, training iterates through the following two steps until convergence:

1. Compute the optimal orthogonal mapping maximizing the similarities for the current dictionaryD:

arg max

WX,WZ

X

i

X

j

Dij((Xi∗WX)·(Zj∗WZ))

An optimal solution is given by WX = U andWZ = V, where U SVT = XTDZ is the singular value decomposition ofXTDZ.

2. Compute the optimal dictionary over the sim-ilarity matrix of the mapped embeddings

XWXWZTZT. This typically uses nearest neighbor retrieval from the source language into the target language, soDij = 1 ifj =

argmaxk(Xi∗WX)·(Zk∗WZ)andDij = 0 otherwise.

The underlying optimization objective is inde-pendent from the initial dictionary, and the algo-rithm is guaranteed to converge to a local opti-mum of it. However, the method does not work if starting from a completely random solution, as it tends to get stuck in poor local optima in that case.

For that reason, we use the unsupervised initial-ization procedure at Section3.2to build an initial solution. However, simply plugging in both meth-ods did not work in our preliminary experiments, as the quality of this initial method is not good enough to avoid poor local optima. For that rea-son, we next propose some key improvements in the dictionary induction step to make self-learning more robust and learn better mappings:

• Stochastic dictionary induction. In or-der to encourage a wior-der exploration of the search space, we make the dictionary induc-tion stochastic by randomly keeping some el-ements in the similarity matrix with probabil-itypand setting the remaining ones to 0. As a consequence, the smaller the value ofpis, the more the induced dictionary will vary from iteration to iteration, thus enabling to escape poor local optima. So as to find a fine-grained solution once the algorithm gets into a good region, we increase this value during train-ing akin to simulated annealtrain-ing, starttrain-ing with

p = 0.1and doubling this value every time

the objective function at step 1 above does not improve more thanǫ = 10−6 for 50 it-erations.

• Frequency-based vocabulary cutoff. The size of the similarity matrix grows quadrat-ically with respect to that of the vocabular-ies. This does not only increase the cost of computing it, but it also makes the number of possible solutions grow exponentially3, pre-sumably making the optimization problem harder. Given that less frequent words can be expected to be noisier, we propose to restrict the dictionary induction process to thekmost frequent words in each language, where we findk= 20,000to work well in practice.

• CSLS retrieval. Dinu et al. (2015) showed that nearest neighbor suffers from the hub-ness problem. This phenomenon is known to occur as an effect of the curse of dimen-sionality, and causes a few points (known as

hubs) to be nearest neighbors of many other points (Radovanovi´c et al.,2010a,b). Among the existing solutions to penalize the similar-ity score of hubs, we adopt the Cross-domain

3There are mn

possible combinations that go from a source vocabulary ofnentries to a target vocabulary ofm

(5)

Similarity Local Scaling (CSLS) from Con-neau et al. (2018). Given two mapped em-beddings x and y, the idea of CSLS is to compute rT(x) and rS(y), the average

co-sine similarity of x and y for their k near-est neighbors in the other language, respec-tively. Having done that, the corrected score

CSLS(x, y) = 2 cos(x, y)rT(x)−rS(y).

Following the authors, we setk= 10.

• Bidirectional dictionary induction. When the dictionary is induced from the source into the target language, not all target language words will be present in it, and some will oc-cur multiple times. We argue that this might accentuate the problem of local optima, as re-peated words might act as strong attractors from which it is difficult to escape. In or-der to mitigate this issue and encourage di-versity, we propose inducing the dictionary in both directions and taking their correspond-ing concatenation, soD=DX→Z+DZ→X.

In order to build theinitial dictionary, we com-puteX′andZ′as detailed in Section3.2and apply the above procedure over them. As the only differ-ence, this first solution does not use the stochastic zeroing in the similarity matrix, as there is no need to encourage diversity (X′ andZ′ are only used once), and the threshold for vocabulary cutoff is set tok= 4,000, soX′andZ′can fit in memory. Having computed the initial dictionary,X′andZ′

are discarded, and the remaining iterations are per-formed over the original embeddingsXandZ.

3.4 Symmetric re-weighting

As part of their multi-step framework, Artetxe et al. (2018a) showed that re-weighting the tar-get language embeddings according to the cross-correlation in each component greatly improved the quality of the induced dictionary. Given the singular value decomposition U SVT = XTDZ, this is equivalent to taking WX = U andWZ =

V S, where X and Z are previously whitened applying the linear transformations (XTX)−1

2 and (ZTZ)−12, and later de-whitened applying

UT(XTX)12U andVT(ZTZ)12V.

However, re-weighting also accentuates the problem of local optima when incorporated into self-learning as, by increasing the relevance of dimensions that best match for the current solu-tion, it discourages to explore other regions of the

search space. For that reason, we propose using it as a final step once self-learning has converged to a good solution. UnlikeArtetxe et al.(2018a), we apply re-weighting symmetrically in both lan-guages, taking WX = U S

1

2 and WZ = V S 1 2. This approach is neutral in the direction of the mapping, and gives good results as shown in our experiments.

4 Experimental settings

Following common practice, we evaluate our method on bilingual lexicon extraction, which measures the accuracy of the induced dictionary in comparison to a gold standard.

As discussed before, previous evaluation has focused on favorable conditions. In particular, ex-isting unsupervised methods have almost exclu-sively been tested on Wikipedia corpora, which is comparable rather than monolingual, exposing a strong cross-lingual signal that is not available in strictly unsupervised settings. In addition to that, some datasets comprise unusually small embed-dings, with only 50 dimensions and around 5,000-10,000 vocabulary items (Zhang et al., 2017a,b). As the only exception, Conneau et al. (2018) re-port positive results on the English-Italian dataset of Dinu et al. (2015) in addition to their main experiments, which are carried out in Wikipedia. While this dataset does use strictly monolingual corpora, it still corresponds to a pair of two rela-tively close indo-european languages.

(6)

challeng-ES-EN IT-EN TR-EN

best avg s t best avg s t best avg s t

Zhang et al.(2017a),λ= 1 71.43 68.18 10 13.2 60.38 56.45 10 12.3 0.00 0.00 0 13.0

Zhang et al.(2017a),λ= 10 70.24 66.37 10 13.0 57.64 52.60 10 12.6 21.07 17.95 10 13.2

Conneau et al.(2018), code 76.18 75.82 10 25.1 67.32 67.00 10 25.9 32.64 14.34 5 25.3

Conneau et al.(2018), paper 76.15 75.81 10 25.1 67.21 60.22 9 25.5 29.79 16.48 7 25.5 Proposed method 76.43 76.28 10 0.6 66.96 66.92 10 0.9 36.10 35.93 10 1.7

Table 1: Results of unsupervised methods on the dataset ofZhang et al.(2017a). We perform 10 runs for each method and report the best and average accuracies (%), the number of successful runs (those with

>5% accuracy) and the average runtime (minutes).

EN-IT EN-DE EN-FI EN-ES

best avg s t best avg s t best avg s t best avg s t

Zhang et al.(2017a),λ= 1 0.00 0.00 0 47.0 0.00 0.00 0 47.0 0.00 0.00 0 45.4 0.00 0.00 0 44.3

Zhang et al.(2017a),λ= 10 0.00 0.00 0 46.6 0.00 0.00 0 46.0 0.07 0.01 0 44.9 0.07 0.01 0 43.0

Conneau et al.(2018), code 45.40 13.55 3 46.1 47.27 42.15 9 45.4 1.62 0.38 0 44.4 36.20 21.23 6 45.3

[image:6.595.119.479.64.148.2]

Conneau et al.(2018), paper 45.27 9.10 2 45.4 0.07 0.01 0 45.0 0.07 0.01 0 44.7 35.47 7.09 2 44.9 Proposed method 48.53 48.13 10 8.9 48.47 48.19 10 7.3 33.50 32.63 10 12.9 37.60 37.33 10 9.1

Table 2: Results of unsupervised methods on the dataset of Dinu et al. (2015) and the extensions of Artetxe et al.(2017,2018a). We perform 10 runs for each method and report the best and average accu-racies (%), the number of successful runs (those with>5% accuracy) and the average runtime (minutes).

ing due to the linguistic distance between them. For completeness, we also test our method in the Spanish-English, Italian-English and Turkish-English datasets of Zhang et al. (2017a), which consist of 50-dimensional CBOW embeddings trained on Wikipedia, as well as gold standard dictionaries4 from Open Multilingual WordNet (Spanish-English and Italian-English) and Google Translate (Turkish-English). The lower dimen-sionality and comparable corpora make an easier scenario, although it also contains a challenging pair of distant languages (Turkish-English).

Our method is implemented in Python using NumPy and CuPy. Together with it, we also test themethodsofZhang et al.(2017a) andConneau et al. (2018) using the publicly available imple-mentations from the authors5. Given that Zhang et al. (2017a) report using a different value of their hyperparameterλfor different language pairs (λ = 10for English-Turkish and λ = 1 for the rest), we test both values in all our experiments to

4The test dictionaries were obtained through personal

communication with the authors. The rest of the language pairs were left out due to licensing issues.

5Despite our efforts,Zhang et al.(2017b) was left out

be-cause: 1) it does not create a one-to-one dictionary, thus diffi-culting direct comparison, 2) it depends on expensive propri-etary software 3) its computational cost is orders of magni-tude higher (running the experiments would have taken sev-eral months).

better understand its effect. In the case ofConneau et al.(2018), we test both the default hyperparam-eters in the source code as well as those reported in the paper, with iterative refinement activated in both cases. Given the instability of these methods, we perform 10 runs for each, and report the best and average accuracies, the number of successful runs (those with >5% accuracy) and the average runtime. All the experiments were run in a single Nvidia Titan Xp.

5 Results and discussion

We first present the main results (§5.1), then the comparison to the state-of-the-art (§5.2), and fi-nally ablation tests to measure the contribution of each component (§5.3).

5.1 Main results

(7)

Supervision Method EN-IT EN-DE EN-FI EN-ES

5k dict.

Mikolov et al.(2013) 34.93† 35.00† 25.91† 27.73†

Faruqui and Dyer(2014) 38.40* 37.13* 27.60* 26.80* Shigeto et al.(2015) 41.53† 43.07† 31.04† 33.73†

Dinu et al.(2015) 37.7 38.93* 29.14* 30.40*

Lazaridou et al.(2015) 40.2 - -

-Xing et al.(2015) 36.87† 41.2728.2331.20

Zhang et al.(2016) 36.73† 40.80† 28.16† 31.07†

Artetxe et al.(2016) 39.27 41.87* 30.62* 31.40*

Artetxe et al.(2017) 39.67 40.87 28.72

-Smith et al.(2017) 43.1 43.33† 29.4235.13

Artetxe et al.(2018a) 45.27 44.13 32.94 36.60

25 dict. Artetxe et al.(2017) 37.27 39.60 28.16

-Init. Smith et al.(2017), cognates 39.9 - - -heurist. Artetxe et al.(2017), num. 39.40 40.27 26.47

-None

Zhang et al.(2017a),λ= 1 0.00* 0.00* 0.00* 0.00* Zhang et al.(2017a),λ= 10 0.00* 0.00* 0.01* 0.01*

Conneau et al.(2018), code‡ 45.15* 46.83* 0.38* 35.38*

Conneau et al.(2018), paper‡ 45.1 0.01* 0.01* 35.44*

[image:7.595.133.464.63.303.2]

Proposed method 48.13 48.19 32.63 37.33

Table 3: Accuracy (%) of the proposed method in comparison with previous work. *Results obtained with the official implementation from the authors. †Results obtained with the framework fromArtetxe et al. (2018a). The remaining results were reported in the original papers. For methods that do not require supervision, we report the average accuracy across 10 runs. ‡For meaningful comparison, runs with<5% accuracy are excluded when computing the average, but note that, unlike ours, their method often gives a degenerated solution (see Table2).

language pairs require different hyperparameters:

λ = 1 works substantially better for Spanish-English and Italian-Spanish-English, but only λ = 10

works for Turkish-English.

The results for the more challenging dataset from Dinu et al. (2015) and the extensions of Artetxe et al. (2017, 2018a) are given in Table 2. In this case, our proposed method obtains the best results in all metrics for all the four language pairs tested. The method ofZhang et al. (2017a) does not work at all in this more challenging sce-nario, which is in line with the negative results re-ported by the authors themselves for similar con-ditions (only %2.53 accuracy in their large Gi-gaword dataset). The method of Conneau et al. (2018) also fails for English-Finnish (only 1.62% in the best run), although it is able to get positive results in some runs for the rest of language pairs. Between the two configurations tested, the default hyperparameters in the code show a more stable behavior.

These results confirm the robustness of the pro-posed method. While the other systems succeed in some runs and fail in others, our method con-verges to a good solution in all runs without

excep-tion and, in fact, it is the only one getting positive results for English-Finnish. In addition to being more robust, our method also obtains substantially better accuracies, surpassing previous methods by at least 1-3 points in all but the easiest pairs. More-over, our method is not sensitive to hyperparame-ters that are difficult to tune without a development set, which is critical in realistic unsupervised con-ditions.

(8)

EN-IT EN-DE EN-FI EN-ES

best avg s t best avg s t best avg s t best avg s t

Full system 48.53 48.13 10 8.9 48.47 48.19 10 7.3 33.50 32.63 10 12.9 37.60 37.33 10 9.1

- Unsup. init. 0.07 0.02 0 16.5 0.00 0.00 0 17.3 0.07 0.01 0 13.8 0.13 0.02 0 15.9

- Stochastic 48.20 48.20 10 2.7 48.13 48.13 10 2.5 0.28 0.28 0 4.3 37.80 37.80 10 2.6 - Cutoff (k=100k) 46.87 46.46 10 114.5 48.27 48.12 10 105.3 31.95 30.78 10 162.5 35.47 34.88 10 185.2 - CSLS 0.00 0.00 0 15.0 0.00 0.00 0 13.8 0.00 0.00 0 13.1 0.00 0.00 0 14.1 - Bidirectional 46.00 45.37 10 5.6 48.27 48.03 10 5.5 31.39 24.86 8 7.8 36.20 35.77 10 7.3

- Re-weighting 46.07 45.61 10 8.4 48.13 47.41 10 7.0 32.94 31.77 10 11.2 36.00 35.45 10 9.1

Table 4: Ablation test on the dataset of Dinu et al. (2015) and the extensions ofArtetxe et al.(2017, 2018a). We perform 10 runs for each method and report the best and average accuracies (%), the number of successful runs (those with>5% accuracy) and the average runtime (minutes).

5.2 Comparison with the state-of-the-art

Table3shows the results of the proposed method in comparison to previous systems, including those with different degrees of supervision. We focus on the widely used English-Italian dataset ofDinu et al. (2015) and its extensions. Despite being fully unsupervised, our method achieves the best results in all language pairs but one, even sur-passing previous supervised approaches. The only exception is English-Finnish, whereArtetxe et al. (2018a) gets marginally better results with a dif-ference of 0.3 points, yet ours is the only unsu-pervised system that works for this pair. At the same time, it is remarkable that the proposed sys-tem gets substantially better results than Artetxe et al.(2017), the only other system based on self-learning, with the additional advantage of being fully unsupervised.

5.3 Ablation test

In order to better understand the role of different aspects in the proposed system, we perform an ab-lation test, where we separately analyze the effect of initialization, the different components of our robust self-learning algorithm, and the final sym-metric weighting. The obtained results are re-ported in Table4.

In concordance with previous work, our results show that self-learning does not work with ran-dom initialization. However, the proposed unsu-pervised initialization is able to overcome this is-sue without the need of any additional informa-tion, performing at par with other character-level heuristics that we tested (e.g. shared numerals).

As for the different self-learning components, we observe that the stochastic dictionary induction is necessary to overcome the problem of poor

lo-cal optima for English-Finnish, although it does not make any difference for the rest of easier lan-guage pairs. The frequency-based vocabulary cut-off also has a positive effect, yielding to slightly better accuracies and much faster runtimes. At the same time, CSLS plays a critical role in the sys-tem, as hubness severely accentuates the problem of local optima in its absence. The bidirectional dictionary induction is also beneficial, contribut-ing to the robustness of the system as shown by English-Finnish and yielding to better accuracies in all cases.

Finally, these results also show that symmet-ric re-weighting contributes positively, bringing an improvement of around 1-2 points without any cost in the execution time.

6 Conclusions

In this paper, we show that previous unsupervised mapping methods (Zhang et al.,2017a;Conneau et al.,2018) often fail on realistic scenarios involv-ing non-comparable corpora and/or distant lan-guages. In contrast to adversarial methods, we propose to use an initial weak mapping that ex-ploits the structure of the embedding spaces in combination with a robust self-learning approach. The results show that our method succeeds in all cases, providing the best results with respect to all previous work on unsupervised and supervised mappings.

(9)

in-termediate vocabularies and re-weighting the fi-nal solution. Our implementation is available as an open source project at https://github. com/artetxem/vecmap.

In the future, we would like to extend the method from the bilingual to the multilingual sce-nario, and go beyond the word level by incorporat-ing embeddincorporat-ings of longer phrases.

Acknowledgments

This research was partially supported by the Spanish MINECO (TUNER TIN2015-65308-C5-1-R, MUSTER PCIN-2015-226 and TADEEP TIN2015-70214-P, cofunded by EU FEDER), the UPV/EHU (excellence research group), and the NVIDIA GPU grant program. Mikel Artetxe en-joys a doctoral grant from the Spanish MECD.

References

Mikel Artetxe, Gorka Labaka, and Eneko Agirre. 2016.

Learning principled bilingual mappings of word em-beddings while preserving monolingual invariance. In Proceedings of the 2016 Conference on Empiri-cal Methods in Natural Language Processing, pages 2289–2294, Austin, Texas. Association for Compu-tational Linguistics.

Mikel Artetxe, Gorka Labaka, and Eneko Agirre. 2017.

Learning bilingual word embeddings with (almost) no bilingual data. InProceedings of the 55th Annual Meeting of the Association for Computational Lin-guistics (Volume 1: Long Papers), pages 451–462, Vancouver, Canada. Association for Computational Linguistics.

Mikel Artetxe, Gorka Labaka, and Eneko Agirre. 2018a. Generalizing and improving bilingual word embedding mappings with a multi-step framework of linear transformations. In Proceedings of the Thirty-Second AAAI Conference on Artificial Intel-ligence (AAAI-18), pages 5012–5019.

Mikel Artetxe, Gorka Labaka, Eneko Agirre, and Kyunghyun Cho. 2018b. Unsupervised neural ma-chine translation. InProceedings of the 6th Inter-national Conference on Learning Representations (ICLR 2018).

Alexis Conneau, Guillaume Lample, Marc’Aurelio Ranzato, Ludovic Denoyer, and Herv´e J´egou. 2018.

Word translation without parallel data. In Proceed-ings of the 6th International Conference on Learning Representations (ICLR 2018).

Georgiana Dinu, Angeliki Lazaridou, and Marco Ba-roni. 2015. Improving zero-shot learning by mitigat-ing the hubness problem. InProceedings of the 3rd International Conference on Learning Representa-tions (ICLR 2015), workshop track.

Manaal Faruqui and Chris Dyer. 2014.Improving vec-tor space word representations using multilingual correlation. InProceedings of the 14th Conference of the European Chapter of the Association for Com-putational Linguistics, pages 462–471, Gothenburg, Sweden. Association for Computational Linguistics.

Guillaume Lample, Alexis Conneau, Ludovic De-noyer, and Marc’Aurelio Ranzato. 2018. Un-supervised machine translation using monolingual corpora only. In Proceedings of the 6th Inter-national Conference on Learning Representations (ICLR 2018).

Angeliki Lazaridou, Georgiana Dinu, and Marco Ba-roni. 2015. Hubness and pollution: Delving into cross-space mapping for zero-shot learning. In Pro-ceedings of the 53rd Annual Meeting of the Associ-ation for ComputAssoci-ational Linguistics and the 7th In-ternational Joint Conference on Natural Language Processing (Volume 1: Long Papers), pages 270– 280, Beijing, China. Association for Computational Linguistics.

Ang Lu, Weiran Wang, Mohit Bansal, Kevin Gimpel, and Karen Livescu. 2015. Deep multilingual cor-relation for improved word embeddings. In Pro-ceedings of the 2015 Conference of the North Amer-ican Chapter of the Association for Computational Linguistics: Human Language Technologies, pages 250–256, Denver, Colorado. Association for Com-putational Linguistics.

Antonio Valerio Miceli Barone. 2016. Towards cross-lingual distributed representations without parallel text trained with adversarial autoencoders. In Pro-ceedings of the 1st Workshop on Representation Learning for NLP, pages 121–126, Berlin, Germany. Association for Computational Linguistics.

Tomas Mikolov, Quoc V Le, and Ilya Sutskever. 2013.

Exploiting similarities among languages for ma-chine translation. arXiv preprint arXiv:1309.4168.

Yves Peirsman and Sebastian Pad´o. 2010. Cross-lingual induction of selectional preferences with bilingual vector spaces. InHuman Language Tech-nologies: The 2010 Annual Conference of the North American Chapter of the Association for Compu-tational Linguistics, pages 921–929, Los Angeles, California. Association for Computational Linguis-tics.

Miloˇs Radovanovi´c, Alexandros Nanopoulos, and Mir-jana Ivanovi´c. 2010a. Hubs in space: Popular near-est neighbors in high-dimensional data. Journal of Machine Learning Research, 11(Sep):2487–2531.

(10)

Yutaro Shigeto, Ikumi Suzuki, Kazuo Hara, Masashi Shimbo, and Yuji Matsumoto. 2015. Ridge regres-sion, hubness, and zero-shot learning. Machine Learning and Knowledge Discovery in Databases: European Conference, ECML PKDD 2015, Pro-ceedings, Part I, pages 135–151.

Samuel L Smith, David HP Turban, Steven Hamblin, and Nils Y Hammerla. 2017. Offline bilingual word vectors, orthogonal transformations and the in-verted softmax. In Proceedings of the 5th Inter-national Conference on Learning Representations (ICLR 2017).

J¨org Tiedemann. 2012. Parallel data, tools and in-terfaces in opus. InProceedings of the Eighth In-ternational Conference on Language Resources and Evaluation (LREC-2012), pages 2214–2218, Istan-bul, Turkey. European Language Resources Associ-ation (ELRA).

Ivan Vuli´c and Anna Korhonen. 2016. On the role of seed lexicons in learning bilingual word embed-dings. InProceedings of the 54th Annual Meeting of the Association for Computational Linguistics (Vol-ume 1: Long Papers), pages 247–257, Berlin, Ger-many. Association for Computational Linguistics.

Ivan Vuli´c and Marie-Francine Moens. 2013. A study on bootstrapping bilingual vector spaces from non-parallel data (and nothing else). In Proceedings of the 2013 Conference on Empirical Methods in Natu-ral Language Processing, pages 1613–1624, Seattle, Washington, USA. Association for Computational Linguistics.

Chao Xing, Dong Wang, Chao Liu, and Yiye Lin. 2015. Normalized word embedding and orthog-onal transform for bilingual word translation. In

Proceedings of the 2015 Conference of the North American Chapter of the Association for Computa-tional Linguistics: Human Language Technologies, pages 1006–1011, Denver, Colorado. Association for Computational Linguistics.

Meng Zhang, Yang Liu, Huanbo Luan, and Maosong Sun. 2017a. Adversarial training for unsupervised bilingual lexicon induction. In Proceedings of the 55th Annual Meeting of the Association for Com-putational Linguistics (Volume 1: Long Papers), pages 1959–1970, Vancouver, Canada. Association for Computational Linguistics.

Meng Zhang, Yang Liu, Huanbo Luan, and Maosong Sun. 2017b. Earth mover’s distance minimization for unsupervised bilingual lexicon induction. In

Proceedings of the 2017 Conference on Empirical Methods in Natural Language Processing, pages 1934–1945, Copenhagen, Denmark. Association for Computational Linguistics.

Yuan Zhang, David Gaddy, Regina Barzilay, and Tommi Jaakkola. 2016. Ten pairs to tag – multilin-gual pos tagging via coarse mapping between em-beddings. InProceedings of the 2016 Conference of

Figure

Figure 1: Motivating example for our unsupervised initialization method, showing the similarity distri-butions of three words (corresponding to the smoothed density estimates from the normalized square rootof the similarity matrices as defined in Section 3.
Table 2: Results of unsupervised methods on the dataset of Dinu et al.racies (%), the number of successful runs (those with (2015) and the extensions ofArtetxe et al
Table 3: Accuracy (%) of the proposed method in comparison with previous work. often gives a degenerated solution (see Tablewithrequire supervision, we report the average accuracy across 10 runs.with the official implementation from the authors.*Results obt

References

Related documents

Other techniques include discussing learning theories and engineering ethics in class, allowing students to use reference sheets for closed-book tests and having students work

Produzione di altri laminati mercantili Produzione di nastri e bande per tubi a caldo Produzione di coils (prodotti finiti).. Produzione di lamiere a caldo di spessore grosso e medio

Here, we demonstrate that heat stability of function (indexed by the Michaelis – Menten constant of the cofactor NADH, K M NADH ) and structure (indexed by rate of denaturation)

The magnitude of the wing-steering response of intact and bulb-ablated flies to a moving figure was about twice as large as the response observed with a visible wide-field stimulus,

Auditory weighting functions were derived from these smoothed functions that may be used to predict perceived levels and correlated noise effects in the harbour porpoise, at least

Microscopy revealed bland epithelium in pools of mucin with focal areas showing papillary tufting, stratification and prominent nucleoli [Table/Fig 2] (Fig

The long NK-target cell conjugation upon killing by lytic granule transfer is distinct from the transient conjugation for FasL-mediated cytotoxicity, which probably ensures

In all animals immediately prior to the onset of emesis, one or more episodes of jaw snapping occurred with the head elevated and the animal raised on its forelimbs.. Each